Heat Transfer in Fixed Beds at Very Low (

Heat Transfer in Fixed Beds at Very Low (...
18 downloads 0 Views 253KB Size
Ind. Eng. Chem. Res. 1997, 36, 3053-3064

3053

Heat Transfer in Fixed Beds at Very Low ( 4, for which the influence of N is low, are plotted as thin-walled symbols. The values for N < 2 are plotted as full symbols, and the values for the intermediate range are plotted as thick-walled symbols. This representation is adhered to in the following Figures 3-11 as well. The kr/kf values plot linearly against Re, and their intercepts are reasonably consistent with the low stagnant thermal conductivity of the packing materials. The magnitudes of the kr/kf and their variation with N is in good agreement with earlier results (Dixon and Yoo, 1992) as will be discussed further below. The Nuw results for spheres shown in Figure 3 are not spread so widely as the kr/kf values, although a similar pattern emerges. For higher N, there is only a small effect of N, then a strong increase at N ) 2.7, a decrease back to values consistent with the high-N results for N ) 2.0, a strong decrease for N ) 1.8, and then a monotonic increase as N decreases from 1.8 to 1.14. The values of Nuw are well-represented by straight lines, although prior work would suggest that Nuw should be proportional to Re0.6-0.75 (Colledge and Paterson, 1984; Dixon and LaBua, 1985; Martin and Nilles, 1993) as Re increases. This apparent linearity may be due to the relatively narrow range of Re achievable in the experimental apparatus of this study. It is interesting that the confidence intervals for Nuw are smaller than those for kr/kf, especially for the lowest N. The reason for this may be attributed to some very flat measured temperature profiles for low N (Dixon, 1994), which also have large temperature “jumps” near the wall, so that hw is well-determined but kr is not. The slopes of the lines fitted to the Nuw data appear to decrease distinctly with decreasing N, and there also seems to be an effect on the intercept. These trends also will be discussed further below. Figures 4 and 5 show the results for full aluminum cylinders. For kr/kf in Figure 4, the values are close in the range 3.5 e N e 6.9, with only N ) 4.7 a little lower than the others. There is then a fairly large decrease for N ) 2.4 and N ) 1.8, both in slope and intercept of the data. The decrease in intercept may be related to the increased void fraction for beds of cylinders in this range (Dixon, 1988a), while the lower values of kr/kf at higher Re may result from bypassing and thus reduced lateral mixing in these beds. Somewhat similar behavior is seen in Figure 5 for Nuw, although there is actually little effect of N until N ) 1.8. An indication of decreasing slope may be inferred from the data. Similar observations may be made in Figures 6 and 7 for packings of aluminum hollow cylinders. In this case, the data are all quite close for kr/kf, with those for N ) 3.5 and N ) 2.4 only a little lower than the others at higher N. For Nuw, all the data are close, except for the results for N ) 5.2, which appear to lie unexpectedly higher than the rest. Note that for these packings it was not possible to reach N < 2, where a stronger influence of N might be expected. All the high conductivity packings in Figures 4-7 give relatively high

3058 Ind. Eng. Chem. Res., Vol. 36, No. 8, 1997

Figure 4. Effect of tube-to-particle diameter ratio N on effective radial conductivity kr for beds of aluminum equilateral cylinders.

Figure 5. Effect of tube-to-particle diameter ratio N on wall Nusselt number Nuw for beds of aluminum equilateral cylinders.

intercepts at Re ) 0, reflecting the effect of the particle conductivity on the stagnant bed contributions to kr/kf and Nuw. In Figures 8 and 9, results for nonporous ceramic hollow cylinders of moderate particle thermal conductivity are presented. In Figure 8, again, the results for kr/kf are close, with only those for N ) 1.8 being significantly lower than the rest. Even this influence

Figure 6. Effect of tube-to-particle diameter ratio N on effective radial conductivity kr for beds of aluminum equilateral hollow cylinders.

Figure 7. Effect of tube-to-particle diameter ratio N on wall Nusselt number Nuw for beds of aluminum equilateral hollow cylinders.

of N is muted for Nuw, and a steady decrease in slope is seen in Figure 9 from N ) 6.9 down to N ) 1.8 with similar intercepts and no sudden drop in magnitude. For the porous ceramic full and hollow cylinders and glass hollow cylinders, the kr/kf values given in Figure 10 show only a minor effect of N, while the results for Nuw in Figure 11 show that N has almost no effect. Only

Ind. Eng. Chem. Res., Vol. 36, No. 8, 1997 3059

Figure 8. Effect of tube-to-particle diameter ratio N on effective radial conductivity kr for beds of solid (nonporous) ceramic hollow cylinders.

Figure 10. Effect of tube-to-particle diameter ratio N on effective radial conductivity kr for beds of porous ceramic full and hollow cylinders and glass hollow cylinders.

Figure 9. Effect of tube-to-particle diameter ratio N on wall Nusselt number Nuw for beds of solid (nonporous) ceramic hollow cylinders.

Figure 11. Effect of tube-to-particle diameter ratio N on wall Nusselt number Nuw for beds of porous ceramic full and hollow cylinders and glass hollow cylinders.

the parameter values for N ) 2.8 and N ) 3.8 are somewhat lower than expected in comparison to those of the other packings, for both kr/kf and Nuw. There is no apparent reason for this observation, except that it may be noted that these packings were both thin-walled glass rings, with larger internal void spaces than those with the other hollow cylinders, all of which had the

internal diameter no larger than 50% of the external diameter. For the glass rings, the ratios were 67% (N ) 3.8) and 75% (N ) 2.8), and both packings had much larger bulk void fractions than those of the other packings. Overall, the data presented in Figures 2-11 confirm and extend the trends reported earlier for packings in

3060 Ind. Eng. Chem. Res., Vol. 36, No. 8, 1997

the range 5 e N e 12 (Dixon, 1988b). The ratio kr/kf is higher for higher conductivity packings and is dependent on particle shape, being higher for full and hollow cylinders than for spheres. The effect of N is minor, except in the very low range, and the data suggest that these are N < 4 for spheres and N < 2 for nonspheres. On the other hand, Nuw is fairly insensitive to particle shape, but does appear to depend systematically on N over the range 1.8 < N < 7, more strongly for spheres than for nonspheres. There is a minor effect of N on the intercept of Nuw when it is correlated linearly against Re. Both kr/kf and Nuw are well-correlated by a linear dependence on Re over the range studied. The data presented in Figures 2-11 were all taken at Re > 100, and the majority were for Re > 200. In this range the high-Re formulas of Dixon and Cresswell (1979) apply, to relate the effective parameters to the underlying individual-phase phenomena, represented by the fluid-phase and solid-phase parameters. These formulas are as follows:

[

]

kr Bis + 4 krs Bif + 4 8 RePr ) + + kf Perf(∞) kf Bif Ns Bis

Figure 12. Comparison of two correlations for limiting fluid-phase radial Peclet number as a function of N to data for spheres from two prior studies.

-1

(7)

and

Nuw )

[

]

Bis + 4 -1 8 krs 8 + + Nuwf + N kf Ns Bis krs 8 Perf(∞) Bis + 4 -1 (8) + Nuwf kf Ns Bis RePr

[

]

written for the two dimensionless groups used in the present study. For reasonably high Re, 8/Ns is small and we may accept the suggestion of Cresswell (1986) to take Bif ≈ Bis to obtain the approximate formula for kr/kf as the sum of a fluid- and a solid-phase contribution

kr krs RePr ) + kf Perf(∞) kf

(9)

in which the solid conductivity term must be recognized as pertaining to the bed center, i.e., it must be estimated from data for which near-wall effects have been accounted for by an apparent wall-to-solid heat transfer coefficient. Similarly, since Nuwf ∝ Re0.6-0.75, the third term on the right-hand side of eq 8 can be neglected for higher Re in comparison to the second to give an approximate formula for Nuw, also as the sum of a fluidphase and a solid-phase contribution

Nuw ) Bis

krs 2 + Nuwf kf N

(10)

where the solid-phase Nusselt number Nuws has been rewritten in terms of a solid-phase Biot number. If the approximate linear formulas of eqs 9 and 10 are applied to the data of Figures 2-11, they may be used to obtain estimates of the parameters Perf(∞), Nuwf, Bis, and krs/kf. These may then be compared to the literature formulas for these parameters to determine whether they extend to the very low range, N < 4. It must be noted that some simplification has been carried out in obtaining eqs 9 and 10, and this and the degree of accuracy in the data indicated by the 95% confidence intervals should be kept in mind when the following results are interpreted.

Figure 13. Comparison of two correlations for limiting fluid-phase radial Peclet number as a function of N to data for full and hollow cylinders.

Fluid-Phase Radial Thermal Conductivity. Values of Perf(∞) were obtained from the slopes of straightline fits to the data, using Pr ) 0.72, and are presented in Figure 12 for the spheres data of Dixon (1994) and in Figure 13 for the nonspheres data of the present work. Also shown in Figure 12 are the values obtained from the spheres data of Dixon and Yoo (1992). Comparisons are made to two frequently-cited formulas for Perf(∞) as a function of N. The first is the early work due to Fahien and Smith (1955)

(

Perf(∞) ) A 1 +

)

19.4 N2

(11)

where in Figure 12 an average A ) 10 was used for spheres and in Figure 13 A ) 5 was used for nonspheres. The second formula is due to Bauer and Schlu¨nder (1978a)

( (

Perf(∞) ) A 2 - 1 -

))

2 2 N

(12)

where A ) 8dpv/xF and the values of xF, which depend on particle shape, are given in the original reference. The data of Figures 12 and 13 clearly show that, for very low N, the strong upturn in Perf(∞) predicted by

Ind. Eng. Chem. Res., Vol. 36, No. 8, 1997 3061

twisting paths for easy fluid bypassing. As N decreases from 2 to 1, the radial displacement to move around a sphere increases, promoting good radial heat transfer and smaller Perf(∞) values. The Bauer and Schlu¨nder (1978a) formula does quite well in predicting Perf(∞) for those packings with high krf, i.e., with good radial displacement, but could seriously overestimate krf for the “in-between” packing states in which there is a high degree of flow bypassing. For nonspheres, there is no strong evidence of any distinguished states for N > 2, and only a couple of data points below N ) 2 indicate any behavior similar to that of spheres. Solid-Phase Radial Thermal Conductivity. The values of kr0/kf obtained from linear extrapolation of the data to zero-flow conditions should represent bed-center values, as the wall heat transfer coefficient was estimated simultaneously. The formula of Bauer and Schlu¨nder (1978b) was derived under similar conditions. Omitting terms relevant to low pressure and surface oxide effects gives

Figure 14. Parity plot for values of static effective radial thermal conductivity predicted by a formula of Bauer and Schlu¨nder (1978b) and values extrapolated from flowing fluid data.

eq 11 is not supported. The Fahien and Smith (1955) formula was, however, developed empirically from data for spheres with N > 6 only, and it is unreasonable to expect it to extrapolate well beyond the conditions for which it was developed. The formula of Bauer and Schlu¨nder (1978a) provides a better representation of the data, especially for nonspheres. The behavior of Perf(∞) for very low N for spheres in Figure 12 is worthy of special attention. As N decreases from 7 to 4, Perf(∞) increases (krf decreases) as expected, due to the increased bypassing of flow along the wall, as the void fraction also increases. At N ) 3, there is an unexpected low value of Perf(∞), followed by higher values between 3 and 2, another low value at N ) 2, high values for N slightly below 2, and then decreasing values as N decreases from 2 toward 1. This pattern is emphasized by the dotted curve in Figure 12. Data from both studies using different tubes are in good agreement, with only a single data point departing from the pattern. An explanation for this behavior may be found by consideration of the packing of spheres in a tube at very low N. At N ) 3, a distinguished packing state can exist, in which the spheres stack in such a way as to “block” any paths for fluid bypassing, thus promoting good radial transport (high krf). The spheres can form a ring packed tightly against the wall, with room for a centerline sphere to sit in the middle of the ring. For 2 < N < 3, the lower values of krf correspond to “holes” in the packing, as the spheres can pack either in a stable torus, leaving a pathway for flow bypassing down the center of the bed, or they can collapse into the bed center, leaving large pathways for flow bypassing at the wall, both arrangements thus reducing the radial displacement of fluid. This phenomenon, and temperature profiles associated with it, were demonstrated previously (Dixon, 1994). At N ) 2, the spheres pack in pairs in alternating layers, thus creating another distinguished state. The centerline hole is closed up at N ) 2, and each layer blocks the large near-wall void spaces of the preceding layer, to promote transverse movement of fluid. This results in the higher krf/lower Perf(∞) values seen. Immediately below N ) 2, the spheres must pack in a staggered single file, with large

kr0 krf0 krs ) + kf kf kf

(

) (1 - x1 - ) 1 +

[

( ) ( )

)

kR 2x1 -  + × kf kf 1- B kp

])

kf B kp kp B+1 B-1 ln + 2 kfB 2 kf kf 1- B 1- B kp kp kf kf -1 x1 -  + (13) kR kp 1-

(

where

dpv kR e ) 2.27 × 10-7 T3 kf 2-e kf

(

)

(14)

and

B)C with

C)

{[

1.25 2.5 2.5 1 +

(1 - )

10/9

( )] di do

2

sphere cylinder

(15)

(16)

hollow cylinder

A comparison between the extrapolated values of the present study and the Bauer/Schlu¨nder formula is presented in Figure 14 in the form of a parity plot. The correlations of Dixon (1988a) were used to predict void fraction for use in the correlation, as these were not measured directly in this work. Overall, there is broad agreement between theory and data, with the effect of particle conductivity being well-represented by the formula. Closer examination, however, reveals that the Bauer/Schlu¨nder formula predicts much less variation with N than is observed in the data. The plotted points tend to lie in horizontal bands, corresponding to significant changes in observed value with N, with little predicted variation with N.

3062 Ind. Eng. Chem. Res., Vol. 36, No. 8, 1997

Figure 15. Comparison of the terms for the effect of N in two correlations for Nuwf with data for spheres.

Figure 16. Comparison of the terms for the effect of N in two correlations for Nuwf with data for full and hollow cylinders.

In deriving their formula, Bauer and Schlu¨nder (1978b) claimed that wall effects on reducing conduction heat transfer would be offset by their effect on increasing radiation heat transfer. Their work was tested for N > 10, in which wall effects would be smaller, and in this case the good agreement they obtained between theory and data would confirm their argument. In the present work for N < 10, wall effects become rather large, and at the present experimental temperatures of 40-100 °C radiation is a quite small contribution, even for large particles. Even allowing for the uncertainties inherent in the extrapolation method applied to the data, we found the present results do not support Bauer and Schlu¨nder’s contention; however, it may be reasonable under higher N and higher temperature conditions. Fluid-Phase Wall Heat Transfer Coefficient. Two formulas for Nuwf have been published in recent years; both were obtained from mass transfer studies and focused on the effects of N. Colledge and Paterson (1984) found

the results are similar to those of Figure 15, and both formulas perform equivalently. The scatter of data seen in Figure 16 should be attributed both to experimental causes and to the use of the approximate eq 10 in fitting the Nuw data. Solid-Phase Wall Heat Transfer Coefficient. The intercepts at zero Re found by fitting eq 10 with Nuwf given by eq 18 give values for Nuw0, which were converted into wall solid Biot numbers using the Bauer/ Schlu¨nder formula for krs/kf to avoid dividing two extrapolated quantities, by the definition Nuw0 ) Bis(krs/kf)(2/N). These values derived from data were then compared to two recent formulas. The first from Martin and Nilles (1993) gives

(

Nuwf ) 0.523 1 -

)

1 Pr1/3Re0.738 ) ACPRe0.738 N

(17)

whereas Dixon and Labua (1985) obtained

(

Nuwf ) 1 -

)

1 Pr1/3Re0.61 ) ADLRe0.61 N

(18)

which are obviously in good agreement as to the influence of N, but differ in the dependence on Re. The present Nuw data have been correlated against Re, Re0.738, and Re0.61. Although differences in slope and intercept were obtained, the differences in the correlation coefficient were so small as to not lead to a preference for any one of the three options. Figure 15 shows the values obtained for ACP with a correlation of Nuw against Re0.738 and those obtained for ADL with a correlation of Nuw against Re0.61. The two formulas both agree with the data for spheres in Figure 15 to about the same extent, making discrimination between them difficult. Note that since both are empirical formulas and are not related to the bedpacking state, neither is able to reproduce the increase seen for N < 2. In Figure 16 similar results are seen with a comparison of the values of ADL obtained by fitting eq 18 to data for nonspheres. Colledge and Paterson (1984) did not develop their correlation for nonspheres so a comparison would not be fair; however,

5 2 Bis ) 1.3 + N N

(19)

while Dixon (1988b) offered

{

2 (2.41 + 0.156(N - 1)2) spheres 2 N Bis ) 2 N (0.48 + 0.192(N - 1)2) nonspheres N

(20)

The two correlations are compared to each other and to data in Figure 17, where it is clear that neither formula is clearly preferred over the other, on the basis of the data range examined here. Indeed, the data exhibit some scatter, and there is almost no dependence on N that can be discerned for the group Bis(2/N). Further investigation will be necessary to decide this point. Conclusions Very strong wall effects have been observed on values of the effective radial thermal conductivity and apparent wall heat transfer coefficient estimated from data obtained in fixed beds of very low N. The high values of kr/kf and Nuw that are found as N approaches unity confirm that a high degree of radial cross-flow is taking place. Systematic and potentially predictable behavior has been found for beds of spheres in the range 1 < N < 2; however, new correlations need to be developed. The possible use of such low-N beds as single-pelletstring reactors will depend on the trade-offs among heat transfer, pressure drop, and particle diffusion limitations.

Ind. Eng. Chem. Res., Vol. 36, No. 8, 1997 3063

Figure 17. Comparison of two correlations for the effect of N on the solid-phase Biot number to data values extrapolated from flowing fluid data.

Current correlations for individual-phase heat transfer mechanisms are acceptable down to N ) 4 for spheres and N ) 2 for nonspheres. More data for Perf(∞) will be needed to elucidate the fine structure for 2 e N e 4 for spheres and to further investigate the region N < 2 for nonspheres. Data for Bis and krs/kf obtained in the present study from extrapolation of flowing fluids results to Re ) 0 were more scattered, and available correlations were in broad agreement with trends in the data. There were some indications that wall effects may influence krs/kf more strongly than current correlations predict, especially over the lower N values and in the absence of significant radiative heat flow. Correlations for Nuwf performed well over the ranges of variables studied. Data on Nuwf over a wide range of Re will be needed to discriminate between correlations with different dependence on Re. In summary, many of the correlations presented by other authors performed unexpectedly well on the current data set, considering that the range of N investigated here is in many cases much lower than that for which the correlations were developed. Acknowledgment The author thanks Amy-Beth Brooks and Darryl Pollica for assistance in the laboratory during the performance of this work. Nomenclature a: specific interfacial surface area Bi: Biot number (hwR/kr) Bif: fluid-phase Biot number (hwfR/krf) Bis: solid-phase Biot number (hwsR/krs) cp: specific heat of gas dpv: diameter of sphere of equivalent volume to particle dt: diameter of tube e: emissivity of the packing G: superficial gas axial flow rate/area h: apparent fluid-to-solid heat transfer coefficient hw: apparent wall heat transfer coefficient hwf: apparent wall-to-fluid heat transfer coefficient hws: apparent wall-to-solid heat transfer coefficient ka: effective axial thermal conductivity

kas: solid-phase effective axial thermal conductivity kf: fluid thermal conductivity kp: particle thermal conductivity kr: effective radial thermal conductivity kr0: effective radial thermal conductivity under stagnant (Re ) 0) conditions krf: fluid-phase effective radial thermal conductivity krs: solid-phase effective radial thermal conductivity kR: radiation conductivity N: tube-to-particle diameter ratio Ns: dimensionless interphase heat transfer coefficient (adt2h/4krs) Nuw: wall Nusselt number (hwdp/kf) Nuwf: wall-to-fluid Nusselt number (hwfdp/kf) Pea: axial Peclet number (Gcpdp/ka) Per: radial Peclet number (Gcpdp/kr) Perf: fluid-phase radial Peclet number (Gcpdp/krf) Perf(∞): limiting Re f ∞ fluid-phase radial Peclet number (Gcpdp/krf) Pr: Prandtl number (µcp/kf) R: radius of tube Re: particle Reynolds number (Gdp/µ) r: radial coordinate T: measured temperature T0: inlet gas temperature to fixed bed Tw: wall temperature of fixed bed v: superficial gas velocity in tube x: dimensionless axial coordinate (z/R) y: dimensionless radial coordinate (r/R) z: axial coordinate Greek Symbols µ: viscosity of gas phase θ: dimensionless temperature ((T - T0)/(Tw - T0)) θwc: dimensionless wall temperature ((Twc - T0)/(Tw - T0))

Literature Cited Ahn, B.; Zoulalian, A.; Smith, J. M. Axial Dispersion in Packed Beds with Large Wall Effects. AIChE J. 1986, 32, 170-174. Bauer, R.; Schu¨nder, E. U. Effective Radial Thermal Conductivity of Packings in Gas Flow. Part I Convective Transport Coefficient. Int. Chem. Eng. 1978a, 18, 181-188. Bauer, R.; Schlu¨nder, E. U. Effective Radial Thermal Conductivity of Packings in Gas Flow. Part II Thermal Conductivity of the Packing Fraction Without Gas Flow. Int. Chem. Eng. 1978b, 18, 189-204. Benenati, R. F.; Brosilow, C. B. Void Fraction Distribution in Beds of Spheres. AIChE J. 1962, 8, 359-361. Borkink, J. G. H.; Borman, P. C.; Westerterp, K. R. Modeling of Radial Heat Transport in Packed Beds - Confidence Intervals of Estimated Parameters and Choice of Boundary Conditions. Chem. Eng. Commun. 1993, 99, 135-155. Chalbi, M.; Castro, J. A.; Rodrigues, A. E.; Zoulalian, A. Heat Transfer Parameters in Fixed Bed Exchangers. Chem. Eng. J. 1987, 34, 89-97. Chu, C. F.; Ng, K. M. Flow in Packed Tubes with a Small Tube to Particle Diameter Ratio. AIChE J. 1989, 35, 148-158. Colledge, R. A.; Paterson, W. R. Heat Transfer at the Wall of a Packed Bed: A j-Factor Analogy Established. In Collected Papers; 11th Annual Research Meeting; Institution of Chemical Engineers: Bath, 1984; 6 pages. Cresswell, D. L. Heat Transfer in Packed Bed Reactors. In Chemical Reactor Design and Technology, NATO Series E No. 110; de Lasa, H. I., Ed.; Martinus Nijhoff: Dordrecht, 1986; pp 687-728. DeWasch, A. P.; Froment, G. F. Heat Transfer in Packed Beds. Chem. Eng. Sci. 1972, 27, 567-576. Dixon, A. G. The Length Effect on Packed Bed Effective Heat Transfer Parameters. Chem. Eng. J. 1985, 31, 163-173. Dixon, A. G. Correlations for Wall and Particle Shape Effects on Fixed Bed Bulk Voidage. Can. J. Chem. Eng. 1988a, 66, 705708. Dixon, A. G. Wall and Particle-Shape Effects on Heat Transfer in Packed Beds. Chem. Eng. Commun. 1988b, 71, 217-237.

3064 Ind. Eng. Chem. Res., Vol. 36, No. 8, 1997 Dixon, A. G. Angular Temperature Variations in Fixed Beds of Spheres. In HTD-236; 29th National Heat Transfer Conference; ASME: Atlanta, 1993; pp 55-64. Dixon, A. G. Heat Transfer in Packed Beds of Spheres with DT/ DP e 4. In Proceedings of the 10th International Heat Transfer Conference; Hemisphere: Brighton, U.K., 1994; Vol. 5, pp 225230. Dixon, A. G.; Cresswell, D. L. Theoretical Prediction of Effective Heat Transfer Parameters in Packed Beds. AIChE J. 1979, 25, 663-676. Dixon, A. G.; LaBua, L. A. Wall-to-Fluid Coefficients for Fixed Bed Heat and Mass Transfer. Int. J. Heat Mass Transfer 1985, 28, 879-881. Dixon, A. G.; Yoo, C. Heat Transfer in Packed Beds of Very Low Tube-to-Particle Diameter Ratio. In Heat and Mass Transfer in Porous Media; Quintard, M., Todorovic, M., Eds.; Elsevier: Amsterdam, 1992; pp 573-584. Fahien, R. W.; Smith, J. M. Mass Transfer in Packed Beds. AIChE J. 1955, 1, 28-37. Govindarao, V. M. H.; Froment, G. F. Voidage Profiles in Packed Beds of Spheres. Chem. Eng. Sci. 1986, 41, 533-539. Govindarao, V. M. H.; Ramrao, K. V. S.; Rao, A. V. S. Structural Characteristics of Packed Beds of Low Aspect Ratio. Chem. Eng. Sci. 1992, 47, 2105-2109. Hsiang, T. C.; Haynes, H. W., Jr. Axial Dispersion in Small Diameter Beds of Large Spherical Particles. Chem. Eng. Sci. 1977, 32, 678-681. Kershenbaum, L. S.; Lopez-Isunza, F. Dynamic Behavior of an Industrial Scale Fixed-Bed Catalytic Reactor. ACS Symp. Ser. 1982, 196, 109-120. Kunii, D.; Suzuki, M.; Ono, N. Heat Transfer from Wall Surface to Packed Beds at High Reynolds Number. J. Chem. Eng. Jpn. 1968, 1, 21-26. Martin, H.; Nilles, M. Radiale Wa¨rmeleitung in Durchstro¨mten Schuttu¨ngsrohren. Chem.-Ing.-Tech. 1993, 65, 1468-1477. McGreavy, C.; Foumeny, E. A.; Javed, K. H. Characterization of Transport Properties for Fixed Beds in Terms of Local Bed Structure and Flow Distribution. Chem. Eng. Sci. 1986, 41, 787-797.

Mongkhonsi, T.; Lopez-Isunza, H. F.; Kershenbaum, L. S. The Distortion of Measured Temperature Profiles in Fixed-Bed Reactors. Trans. Inst. Chem. Eng. 1992, 70A, 255-264. Mueller, G. E. Radial Void Fraction Distributions in Randomly Packed Fixed Beds of Uniformly Sized Spheres in Cylindrical Containers. Powder Technol. 1992, 72, 269-275. Mueller, G. E. Angular Void Fraction Distributions in Randomly Packed Fixed Beds of Uniformly Sized Spheres in Cylindrical Containers. Powder Technol. 1993, 77, 313-319. Olbrich, W. E.; Potter, O. E. Heat Transfer in Small Diameter Packed Beds. Chem. Eng. Sci. 1972a, 27, 1723-1732. Olbrich, W. E.; Potter, O. E. Mass Transfer from the Wall in Small Diameter Packed Beds. Chem. Eng. Sci. 1972b, 27, 1733-1743. Rao, S. M.; Toor, H. L. Heat Transfer Between Particles in Packed Beds. Ind. Eng. Chem. Fundam. 1984, 23, 294-298. Scott, D. S.; Lee, W.; Papa, J. The Measurement of Transport Coefficients in Gas-Solid Heterogeneous Reactions. Chem. Eng. Sci. 1974, 29, 2155-2167. Solcova, O.; Schneider, P. Axial Dispersion in Single-Pellet-String Columns Packed with Cylindrical Particles. Chem. Eng. Sci. 1994, 49, 401-408. Vortmeyer, D.; Winter, R. P. On the Validity Limits of Packed Bed Reactor Continuum Models with Respect to Tube to Particle Diameter Ratio. Chem. Eng. Sci. 1984, 39, 1430-1432.

Received for review September 30, 1996 Revised manuscript received February 5, 1997 Accepted February 7, 1997X IE9605950

Abstract published in Advance ACS Abstracts, June 15, 1997. X